CH 10

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

C&S in Pastures Chap 10 22/10/01 1:39 pm Page 193

10 Competition and Environmental Stress in


Temperate Grasslands

Duane A. Peltzer* and Scott D. Wilson


Department of Biology, University of Regina, Regina, Saskatchewan, Canada

Introduction 1987; Aerts and van der Peijl, 1993; Berendse,


1994). It is not clear, however, whether these traits
There is abundant evidence that competition helps identify good stress tolerators or good competitors
to determine the species composition of temperate in stressful environments (Grime, 1977; Huston
grasslands (Harper, 1977; Lauenroth and Aguilera, and Smith, 1987; Tilman, 1988; Berendse and
1998; Wilson, 1998, 1999). One approach to Elberse, 1990). Whereas many studies have examined
organizing knowledge about competition is to competitive interactions along natural and experi-
examine how it varies along environmental mental productivity gradients in grasslands (Wilson
gradients. Here we review which traits confer and Tilman, 1991, 1993, 1995; Reader and Bonser,
competitive ability to plants and the role of compe- 1993; Reader et al., 1994; Peltzer et al., 1998; see also
tition along environmental gradients of stress. reviews by Goldberg and Barton, 1992; Gurevitch et
Plant competitive ability can be divided into two al., 1992), relatively few studies have explored how
components: competitive response and competitive suites of traits responsible for stress tolerance affect
effect (Goldberg, 1990). Good response competitors competitive ability, or which traits should be related
are those species able to resist suppression by competi- to competitive ability along environmental gradients.
tors. On the other hand, good effect competitors are How does the role of competition vary along gra-
able to reduce the performance of other species. Traits dients of environmental stress? Here we follow
conferring either aspect of competitive ability, Grime’s (1979) definition of stress as any factor that
response or effect, may change along stress gradients. limits biomass production. In grasslands, biomass is
For example, plants from unproductive, stressful envi- typically limited by soil resources: water (Lauenroth
ronments tend to have high below-ground biomass et al., 1978; Sala et al., 1988; Silvertown et al., 1994)
allocation, long-lived tissues, high nutrient retention or nutrients (Tilman, 1987; Berendse, 1994). Peak
and high tolerance of water and nutrient stress standing crop (i.e. above-ground plant biomass) is
(Chapin, 1980; Chapin et al., 1993); these traits may frequently used as a measure of stress in herbaceous
allow them to displace plants from more productive grassland communities: low standing crop reflects
environments (see review in Goldberg, 1990). high stress and low soil resources. Grime (1979) pre-
Both field and garden experiments show that dicted that stress should limit plant growth rates and
stress-tolerant species can dominate vegetation over resource demand. Thus, competition might be rela-
the long term, due to higher levels of nutrient tively unimportant on dry or nutrient-poor soils that
retention and tissue longevity (Berendse and Aerts, support little standing crop. Competition might

*Present address: Landcare Research, Lincoln, New Zealand.


© CAB International 2001. Competition and Succession in Pastures
(eds P.G. Tow and A. Lazenby) 193
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 194

194 D.A. Peltzer and S.D. Wilson

increase in importance as soil resources increase and (competitive response), and the ability of plants to
standing crop increases. This idea can be traced back reduce the performance of other species (competitive
to Darwin (Harper, 1977) and is still current effect). Traits conferring success for both response
(Keddy, 1989). Alternatively, stressful habitats may and effect may change along stress gradients.
be characterized by intense competition for the
resources that limit growth, such as nutrients and
water (Newman, 1973; Tilman, 1988). As resources Competitive responses
increase, stress decreases, standing crop increases,
shade increases and competition may shift from Competitive response ability (sensu Goldberg,
being mostly below ground to mostly above ground. 1990) is the ability of a plant to resist suppression
We review evidence for the hypothesis that the by neighbours (i.e. the resident vegetation). This
role of competition in controlling community can be measured as the change in performance of a
structure changes along gradients of stress, resource target plant or species in response to the presence of
availability and standing crop. Our focus is on field neighbours, either conspecific or interspecific.
experiments that examine the role of competition There are several reasons for using small target
and variation in competitive ability along gradients plants in existing vegetation (Goldberg, 1990).
of stress in natural temperate grasslands. We First, all plants must regenerate and pass through
include mid-latitude arid and alpine grasslands but critical early life stages in order to establish; early
not tropical grasslands or savannahs or Arctic life stages may be more sensitive to the effects of
graminoid communities. neighbours and the environment than established
adult plants (Grubb, 1977). Thus, individual perfor-
Competitive Ability mance (e.g. seedling survival and growth) can be
linked with population processes (e.g. recruitment,
The relationship between competitive ability and the population growth, distribution). Secondly, competi-
distribution or abundance of species in communities tive response ability determines which species persist
is still poorly understood, despite many experiments in a habitat to contribute to community-level diversity
on competitive interactions in the field (see reviews and productivity (Wilson and Tilman, 1995). Thirdly,
by Goldberg and Barton, 1992; Gurevitch et al., good response competitors are able to withstand
1992; Goldberg, 1996). Several important questions resource shortages imposed by competing plants and
remain. What is competitive ability? Does competi- are also likely to persist in stressful habitats.
tive ability vary among species or environments? Plant traits associated with competitive response
How important is competitive ability relative to abi- ability are those which allow a species to persist and
otic factors, other ecological processes, and historical perform well in the presence of neighbours, and
factors (Givnish, 1986; Welden and Slauson, 1986; thus resource shortages or stress. Similarly, the best
Felsenstein, 1988; Keddy, 1989; Harvey and Pagel, competitors in Tilman’s (1982) resource ratio
1991; Aarssen, 1992; Underwood and Petraitis, hypothesis model of plant competition are those
1993; Westoby et al., 1995; Silvertown and Dodd, species having the lowest resource requirement, or
1996)? Much interest has focused on whether the R*. Empirical support for this model was provided
abundance and distribution of species are the result by Wedin and Tilman’s (1993) experimental mix-
of variation in stress tolerance, competitive ability or tures of native prairie grasses. In their study, grasses
niche differentiation among species (Grime, 1977; with the lowest R* values won in competition,
Grubb, 1977; Chapin, 1980; Keddy, 1989; Smith regardless of initial planting densities in mixtures.
and Huston, 1989; Austin, 1990; Aarssen, 1992; Generally, traits suggested to confer competitive
McLellan et al., 1997). Here we discuss the relation- response ability in relatively unproductive systems
ships among competitive ability, stress tolerance and are identical to those suggested for stress tolerance
community composition. (Table 10.1). Typically, these species have high
root : shoot ratios and low growth rates, are small
and have nutrient-conserving mechanisms, such as
Components of competitive ability long-lived tissues, carbon-based defences and
storage organs (Grime, 1977, 1979; Chapin, 1980,
Goldberg (1990) distinguishes between the ability of 1991; Tilman, 1988; Berendse and Elberse, 1990;
plants to perform well in the presence of neighbours Chapin et al., 1990).
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 195

Competition and Environmental Stress 195

Table 10.1. Summary of plant traits suggested to confer high competitive response or competitive effect
ability at low and high primary productivity. The last column shows traits associated with stress tolerance.
Traits for competitive ability are modified from Goldberg (1996).

Response Effect

Low High Low High


productivity productivity productivity productivity Tolerance

Leaf allocation Lowabf Lowaf Low High Lowde


Stem allocation Lowab Higha Low High Lowde
Root allocation Highabf Lowaf High Low Highde
Reproductive allocation Lowab Lowa High Lowcg Lowcdi
Growth rate Lowabf Lowabf High Highc Lowcde
Litter production Low High High Highc Lowc
Photosynthetic rate Lowa High High High Lowe
Height at maturity Lowa Higha High Highc Lowc
Leaf area Low High High Highg Lowci
Plant mass Low Higha High Highc Low
Specific root length Higha High High Highh High
Leaf area/mass Lowaf Highaf Low Highcg Low
Tissue longevity Highag Lowadg High Lowcg Highcdei
Plasticity Low Low High Highc Lowcd
Tissue [N] Lowabg Highbg High High Lowde
Rate of nutrient uptake Lowb Lowb High Highc Lowe
Nutrient storage High Low High Both Highdi
Nutrient losses or leaching Lowb Highb High High Lowd
Nutrient foraging ability High Low High Both Low
Nutrient-based defences Highb Lowb Low Lowc Highe
C-based defences Highe Low Low Low High
Mycorrhizal infection Highg Low Low Low Highd
Shade tolerance Lowbf Highaf Low Lowc Low
Drought tolerance Highf Lowf Low Lowc High
Low-nutrient tolerance Highbg Lowag Low Lowc High

a, Tilman (1988); b, Tilman (1990); c, Grime (1977); d, Chapin (1980); e, Chapin et al. (1993); f, Smith
and Huston (1989); g, Berendse and Elberse (1990); h, Caldwell and Richards (1986); i, Grime (1979).

Why might variation in competitive response be (1995) measured the competitive responses of eight
important in grasslands? Variation in competitive grassland species (four forbs and four grasses) in an
responses among species may determine species old field in Minnesota. The species showing the
positions within a competitive hierarchy. This in weakest response, i.e. the species most suppressed by
turn may determine their distributions along stress neighbours, was the numerical dominant of natural
gradients or their relative abundance in a community vegetation, the perennial grass Schizachyrium scopar-
(Grime, 1979; Keddy, 1989). The relationship ium. This result was robust across four combinations
between species competitive responses and distribu- of nitrogen (N) availability and soil disturbance.
tions is relatively well documented in wetlands (e.g.
Wilson and Keddy, 1986; Gaudet and Keddy, 1988; STRESS TOLERANCE AND COMPETITIVE RESPONSES.
Keddy and Shipley, 1989) but less well understood Because competition is often for resources and
in grasslands (Mitchley and Grubb, 1986; Aarssen, many of the traits associated with stress tolerance
1988; Wilson and Tilman, 1995). Generally, there is interact with patterns of resource availability, there
no strong relationship between competitive should be a close link between competitive response
responses and the distribution or abundance of ability and stress tolerance. This is not a new idea;
species along stress gradients (Herben and Krahulec, Grime (1977) listed traits associated with stress tol-
1990; Silvertown and Dale, 1991; Shipley and erators and competitive species two decades ago. For
Keddy, 1993). For example, Wilson and Tilman Grime (1977), any energy a plant spends coping
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 196

196 D.A. Peltzer and S.D. Wilson

with stress decreases competitive ability, so that spaced ramets (‘guerilla’ strategy) were more
there is a trade-off between stress tolerance and strongly suppressed by L. perenne than genotypes
competitive ability. In contrast, other authors define producing few, closely spaced ramets (‘phalanx’
a good competitor as a species able to perform well strategy). The role of within-species variation in
despite resource shortages (Tilman, 1982, 1988; competitive response in determining species per-
Berendse and Elberse, 1990; reviewed by Grace sistence in natural communities deserves further
1990, 1991). attention.
The distinction is this: species in productive
environments are competitive if they exploit RESPONSES TO HETEROGENEITY. Species may respond
resources as quickly and efficiently as possible differently to resource heterogeneity. Such hetero-
(Grime, 1977). Resource competition may also be geneity may occur in both space (Campbell and
intense in unproductive habitats (Wilson and Grime, 1989; Campbell et al., 1991; Caldwell and
Tilman, 1991, 1993, 1995; Wilson, 1993a, b), but Pearcy, 1994; Casper and Cahill, 1998) and time
the species that lower limiting resources to the (Fitter, 1986; Campbell and Grime, 1989;
lowest level and use them most efficiently may Bilbrough and Caldwell, 1997; Goldberg and
dominate the vegetation (Tilman, 1982). Thus, Novoplansky, 1997).
there may be no trade-off between stress tolerance Recent studies have shown that plants respond
and competitive ability. Stress tolerance and com- to soil nutrient patchiness independently from
petitive ability can be conferred by the same traits resource level (Hutchings, 1988; Casper and Cahill,
(Tilman, 1982). Traits predicted to confer both 1998). There is abundant evidence that the ability
competitive response ability and stress tolerance are to forage for patchy resources differs among species
summarized in Table 10.1. of grasses (Jackson et al., 1990; Humphrey and
Pyke, 1997; Reynolds et al., 1997). Further, there
WITHIN-SPECIES VARIATION IN COMPETITIVE may be a trade-off between the precision of
RESPONSES. Within-species or within-population foraging and the size of resource patches exploited
variability in competitive ability may also con- (Campbell et al., 1991). Large, rhizomatous grasses
tribute to species coexistence and community should be better foragers than tussock grasses, but
composition. Within-species variability may dimin- their foraging precision should be low, due to wider
ish among-species differences in competitive ability. spacing between ramets.
Genotypic variability and specificity of interactions Species may also differ in their responses to
with neighbouring plants or among sites have been temporal heterogeneity (Grime, 1979; Chapin,
shown in pastures and old fields (Aarssen, 1988; 1980; Goldberg and Novoplansky, 1997). In nat-
Mehrhoff and Turkington, 1990; Turkington, ural grasslands, such as the North American Great
1991). For example, Mehrhoff and Turkington Plains, soil resources and water may be available
(1990) compared the competitive ability of five only during infrequent pulses (Sala and Lauenroth,
populations of Trifolium repens from different-aged 1985; Sala et al., 1992), and differences in ability
pastures. They found considerable variation in to exploit temporal patches may contribute to
competitive ability among populations of T. repens competitive success (Caldwell, 1994). Campbell
in pot, garden and pasture competition experi- and Grime (1989) showed that a relatively slow-
ments with grasses. growing grass, Festuca ovina, was able to use short
Aarssen (1992) reviewed several theories nutrient pulses of from 0.1 to 10 h in duration,
regarding variation in competitive abilities at the while a faster-growing grass, Arrhenatherum elatius,
genotype level; his working hypothesis is that varia- could use only longer pulses. Grime (1994) sug-
tion in competitive ability among genotypes within gests that species occurring in habitats with ‘chronic
a species is as great as variation among species. One nutrient stress’ have large, long-lived root systems,
implication of this hypothesis is that competitive which remain functional throughout the year and
exclusion will not occur at the level of the species in are capable of utilizing short resource pulses. In
a community. Cheplick (1997) explored within- contrast, faster-growing species tend to produce
species variation in competitive response using 11 new tissues to capture additional resources, making
genotypes of the rhizomatous perennial grass their responses to pulses much slower. Although
Amphibromus scabrivalis grown in competition with resource allocation has received much attention,
Lolium perenne. Genotypes with more widely grasses may also vary in their physiological ability
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 197

Competition and Environmental Stress 197

to capture nutrients in response to resource pulses tremuloides) forest. Over one growing season,
(Mouat, 1983; Caldwell, 1994). Interest in the prairie grasses significantly lowered the heterogene-
importance of plant responses to resource hetero- ity of both available N and moisture. Thus, grasses
geneity is increasing (Shorrocks and Swingland, decreased the patchiness of soil resources.
1990; Bell and Lechowicz, 1994; Caldwell and
Pearcy, 1994; Grime, 1994; Miller et al., 1995; STRESS AND COMPETITIVE EFFECTS. Stress is the result
Bilbrough and Caldwell, 1997; Casper and Cahill, of both biotic (due to competitive effects) and abi-
1998), but much remains to be done in grasslands. otic (environmental) processes (Goldberg and
Novoplansky, 1997). Interactions between biotic
and abiotic stress may occur if competitive effect
Competitive effects
varies with environment. Few studies to date have
Competitive effect ability (sensu Goldberg, 1990) is separated stress caused by neighbouring vegetation
defined as the ability of a neighbouring plant or from stress imposed by the environment. Such stud-
species to suppress a focal individual or species. ies are needed to understand the relationship
Competitive effects are quantified as the per-plant between biotic and abiotic stresses in grasslands.
or per-unit-mass (e.g. per-gram) reduction in per- Austin (1990) suggests that environmental influ-
formance of the focal plant or species (see review by ences on species performance are as important as
Goldberg and Scheiner, 1993). Competitive effects species interactions. It is clear that both influence
are often assumed to be equivalent among neigh- grasses. For example, in an old field in Minnesota,
bour species, because all plants use the same increasing N availability increased the growth of
resources (e.g. light, water, nutrients) and these transplanted grasses by about 90%; removing neigh-
resources are supplied along gradients rather than as bours increased growth by 40%; and both increas-
discrete packages (Harper, 1965; Goldberg and ing N and removing neighbours increased growth
Werner, 1983; Goldberg, 1996). As noted above, by 150% (see Fig. 1 in Wilson and Tilman, 1995).
this assumption requires more investigation. Bakker Some questions remain. Are species with larger
(1996) found differences in competitive effects competitive effects less tolerant of harsh or fluctuat-
between introduced and native grasses in mixed- ing abiotic conditions? The distinction between
grass prairie. Wedin and Tilman (1993) also found biotic and abiotic stress is important because they
significant differences between early- and late- probably act at different scales. Abiotic stress
successional grasses; later grasses lowered soil N gradients may occur at larger scales, influencing
availability more than early grasses, which may entire communities, and filtering the local species
explain their eventual dominance of old fields. pool by removing those species that are physiologi-
Competitive effects may occur at several scales. cally incapable of tolerating local environmental
In the short term, plants reduce resource levels and conditions (Harper, 1977). Variation in competi-
create localized zones of resource depletion tive effect ability, and thus biotic stress, acts to
(Caldwell, 1994; Huston and DeAngelis, 1994). eliminate species after environmental filtering
Over longer time periods, plants may modify through competitive exclusion, similar to
nutrient cycles through shoot or root litter quality Diamond’s (1975) assembly rules. This model has
or plant–soil community feedbacks; other long- been applied to wetlands (Keddy, 1992), but has
term feedbacks may arise through mutualisms, such not been tested explicitly in grasslands. Part of the
as mycorrhizas, or through diseases (Allen and reason may be that there are few contrasting envi-
Allen, 1990; Fitter, 1991; Read, 1991; Wilson and ronments with discrete boundaries in grasslands;
Agnew, 1992; Wedin and Pastor, 1993; Bever, instead, both biotic and abiotic stresses commonly
1994; Stark, 1994; Bever et al., 1997; Kleb and occur along gradients.
Wilson, 1997; Wilson, 1998).
In summary, good response competitors are able to
EFFECTS ON HETEROGENEITY. Plants can also influ- withstand resource shortages imposed by neigh-
ence the spatial heterogeneity of soil resources, bours. On the other hand, good effect competitors
water and light in grasslands (Hook et al., 1991; are able to strongly suppress the performance of
Caldwell, 1994). For example, Kleb and Wilson competing species. However, these two components
(1997) used a reciprocal soil transplant experiment of competitive ability are not necessarily correlated
between mixed-grass prairie and aspen (Populus (Goldberg, 1990). One way to compare competitive
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 198

198 D.A. Peltzer and S.D. Wilson

ability among species or along gradients of stress is Does Competition Vary with
by measuring competition intensity, the relative
decline in species performance caused by competi- Standing Crop?
tion.
Stress is a common phenomenon in natural grass-
lands. Plant productivity, usually measured as
aboveground biomass, is normally limited by soil
Measuring Competition resources: water and nutrients (Lauenroth et al.,
in the Field 1978; Tilman, 1987; Sala et al., 1988; Silvertown et
al., 1994). Stress can occur along natural gradients
A simple way to measure the intensity of competi- – for example, rainfall gradients or north- vs.
tion in the field is to compare the performance of south-facing slopes. Stress can also vary through
transplants in plots without neighbours with that in time with among-year variation in precipitation
plots with neighbours. Any difference in perfor- and temperature. Here we review how competi-
mance is presumably attributable to the presence of tion varies along natural and experimental stress
neighbours. This can be done at sites differing in gradients.
some other factor of interest, e.g. stress or grazing
intensity. It can also be done with different trans-
plant species in order to compare their responses. Is standing crop a reliable indicator
Comparisons of competition intensity (CI) among of stress?
sites or species should be standardized for the effect
of the sites or species identities on transplant per- Standing crop should reflect stress from a physio-
formance (Wilson and Keddy, 1986; Grace, 1995; logical point of view, but other biotic factors may
Miller, 1996) as: influence standing crop, making it a poor indicator
CI = (NN – AN) / NN of stress. Standing crop, of course, reflects rates of
biomass removal, as well as production. Thus low
where NN is transplant performance (e.g. survivor- standing crop sites may be produced by high dis-
ship, mass, growth, seed production, tiller number) turbance rates. Standing crop in a Minnesota old-
in plots with no neighbours present and AN is field grassland was enhanced with three levels of N
performance with all neighbours present. addition; controls made a fourth treatment. All N
Regression analysis can be used to examine levels were crossed with four levels of soil distur-
relationships between CI and stress or standing bance applied with a mechanical tiller, which
crop (e.g. Wilson and Keddy, 1986; Belcher et al., produced 0, 25, 50 and 100% bare ground at the
1995; Grace, 1995; Miller, 1996; Peltzer et al., start of each growing season. Large mammals and
1998). Care should be taken, however, not to use abiotic disturbance (e.g. fire) were excluded from
transplant performance as a measure of stress: in the experiment, so that the effects of N availability
this case, the same term (e.g. growth in the absence and disturbance could be studied independently.
of neighbours) ends up on both sides of the regres- Competition intensity, measured by transplants of
sion equation, producing a spurious correlation. the grass S. scoparium (Wilson and Tilman, 1993)
Alternatively, such experiments can be examined and, in a later experiment, using transplants of
with analysis of variance (ANOVA) of transplant eight species (Wilson and Tilman, 1995), decreased
performance: a significant interaction between significantly with increasing disturbance, but did
competition and site treatments would suggests not vary with N availability. The results suggest
that competition intensity varies among sites (e.g. that low- standing- crop habitats may have low
Platenkamp and Foin, 1990; Goldberg levels of competition if the low standing crop is
and Scheiner, 1993; Wilson and Tilman, 1995). partly attributable to disturbance. Removal of
ANOVA of performance revealing significant biomass by disturbance decreases community
interactions between transplant species and demand on resources and decreases competition
competition suggest interspecific differences in (Taylor et al., 1990). In a series of studies, Reader
competitive ability; three-way interactions, includ- (1992, 1993) and Reader and Bonser (1993) per-
ing site, suggest that these differences vary with formed plant removal experiments (i.e. removing
stress (Wilson and Tilman, 1995). one or more species from intact vegetation (sensu
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 199

Competition and Environmental Stress 199

Keddy, 1989) at varying levels of stress and her- Bonser, 1993), but not for P. compressa, which was
bivory. Competition affected transplants only in equally suppressed in both habitats. Bonser and
plots caged to exclude grazers. In open plots, Reader (1995) used a wider selection of standing-
grazers were far more important than competition crop values and found competition intensity to
in controlling plant success. increase with standing crop. In Montana, grass
neighbours inhibited an annual mustard in a wet
year but facilitated it in a dry year (Greenlee and
Natural stress gradients Callaway, 1996).

In natural vegetation, variation in standing crop is


often inferred to reflect variation in stress. Reader et Experimental stress gradients
al. (1994) grew seedlings of Poa pratensis in cleared
plots and intact vegetation in fields in Australia, Other transplant experiments have been performed
Europe and North America. Competition intensity along experimental fertility gradients. As in the case
did not increase with standing crop when all sites of natural gradients, they produce conflicting
were examined together. Transplant experiments results about variation in competition intensity.
along a gradient of soil depth and standing crop in Reader and Best’s (1989) Hieracium experiment
Ontario also found no variation in competition described above found significant variation in com-
intensity (Belcher et al., 1995). Similar results were petition intensity associated with natural variation
found in an old field in Michigan (Foster and in standing crop, but no variation in competition
Gross, 1997); in this case, however, competition intensity was produced by supplying extra water to
intensity increased significantly with neighbour- the vegetation. Wilson and Shay (1990) removed
hood litter mass. A meta-analysis of 34 studies neighbours from around established grass tussocks
found no difference in competition intensity in Manitoba mixed-grass prairie and found that
between more stressful habitats (desert and Arctic) neighbour removal increased tussock size to a similar
and less stressful habitats (prairies, meadows and extent in both fertilized and unfertilized plots.
old fields) (Gurevitch et al., 1992). Burning also had no effect on competition intensity.
Other experiments, in contrast, have found Wilson and Tilman (1991) transplanted three grass
variation in competition intensity. Del Moral species into clearings and intact vegetation within a
(1983) grew transplants in two alpine grasslands in 5-year-old N addition experiment in Minnesota:
Washington. In one neighbourhood, with 50 g neighbours suppressed transplant growth to the
m2 standing crop, neighbours increased transplant same extent at all N levels. Similar experiments in a
survival. In a second neighbourhood, with 650 g nearby field with one (Wilson and Tilman, 1993),
m2 standing crop, no transplants survived in the two (Wilson, 1994) and eight species (Wilson and
presence of neighbours. This result suggests that Tilman, 1995) produced the same result. Further,
neighbours were facilitative at low standing crop competition did not increase with added N in
but competitive at high. Gurevitch (1986) followed either undisturbed or tilled plots, indicating that
the fate of naturally establishing seedlings of grasses both perennial and annual grass neighbourhoods
along a topographic gradient in Arizona. had similar behaviour along the experimental fertil-
Neighbours decreased seedling performance more ity gradient (Wilson and Tilman, 1993, 1995).
in moister, low-lying sites than on dry ridge tops. DiTommaso and Aarssen (1991) transplanted three
Reader and Best (1989) removed neighbours in an grass species into clearings and intact vegetation in
old field in Ontario and followed the population fertilized and unfertilized plots in an old field in
response of the composite Hieracium floribundum Ontario and found that fertility had no effect on
in low- and high-standing-crop sites. Initial competition intensity. Grass transplants grown for
Hieracium densities were made similar by thinning 3 years in a 5-year-old N addition experiment in an
all plots at the start of the experiment. old field in Saskatchewan were equally suppressed
Neighbours reduced Hieracium performance at at all levels of N availability (Peltzer et al., 1998).
high-standing-crop sites but had no effect at low Tree and shrub transplants in the same field were
standing crop. Similar results were found for three equally suppressed, regardless of whether the
species of transplants in the same system (Reader, vegetation was supplied with extra water and N (Li
1992) and for the grass Poa pratensis (Reader and and Wilson, 1998).
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 200

200 D.A. Peltzer and S.D. Wilson

One exception to the trend of competition 1996, personal communication). Competition


intensity not varying along experimental gradients intensity (CI) was calculated as above for each
is given by Reader (1990), who examined the species, year and vegetation type.
impact of neighbours on H. floribundum in There was no clear relationship between compe-
unfertilized and fertilized plots: neighbour removal tition intensity and standing crop (Fig. 10.1), either
increased Hieracium recruitment and survival only for the complete data set or for the data set with the
in fertilized plots. large study of Reader et al. (1994) excluded. Thus,
our summary of other studies corroborates the lack
of an obvious pattern in competition intensity
Summary of competition intensity and found by Reader et al. (1994). Although there is
standing crop always a danger of falsely failing to reject the null
hypothesis of no effect, an analysis of the same data
With one exception, competition intensity did not set, but including recently disturbed plots, found
vary with stress along experimental gradients. that competition intensity did increase significantly
Competition intensity frequently increased with with field age (Wilson, 1999). This disparity
standing crop along natural gradients, but not suggests that standing crop is a relatively poor
always. In some cases, the increase in competition predictor of competition intensity. Habitats which
intensity was produced by a single point where are so stressful that plants cannot maintain live
neighbours had no negative effect at low standing mass are, of course, likely to have little competition
crop (Del Moral, 1983; Bonser and Reader, 1995; (e.g. Grubb, 1992; Belcher et al., 1995; Kadmon,
Greenlee and Callaway, 1996). Overall, the results 1995; Goldberg and Novoplansky, 1997), but
support the idea that competition intensity this description probably does not apply to tem-
increases with standing crop on natural gradients perate grasslands.
more than on experimental gradients (Goldberg Lastly, dominant species may differ in their
and Barton, 1992). competitive effects regardless of stress, so that in
The results of any experiment depend on the any particular environment, the intensity of
conditions under which it is carried out, and competition is determined by the identity of the
measurements of competition intensity have been dominant species and not by stress or resource
carried out within a wide range of standing crop availability. Bakker (1996) grew transplants with
values (summarized in Belcher et al., 1995). We and without neighbours in two sections of a 50-
summarized the results of several studies that gave year-old Saskatchewan field. One section had been
the growth rate of grass transplants with and with- planted with the introduced pasture grass Agropyron
out herbaceous neighbours. Some studies reported cristatum, and the other section had undergone
final transplant mass (Wilson, 1994; Gerry and natural succession to native prairie grasses.
Wilson, 1995) but were included because they Competition intensity varied little between
reported initial mass and allowed calculation of dominant vegetation types, but the F ratio for the
growth rates. The studies were performed in fields competition term in the ANOVA of transplant
abandoned from cultivation for at least 10 years. growth was twice as high in the Agropyron-
We included studies in which both neighbour roots dominated vegetation than in the native-
and shoots were removed and for which standing dominated vegetation, suggesting that Agropyron
crop could be determined (Wilson and Tilman, exerted greater competitive effects than did native
1991, 1993, 1995; Wilson, 1993a, 1994; Reader et species (Underwood and Petraitis, 1993). Further,
al., 1994; Gerry and Wilson, 1995; Bakker, 1996; the growth of transplants of the native grass
Foster and Gross, 1997; Peltzer et al., 1998; S.D. Bouteloua gracilis decreased significantly with neigh-
Wilson, unpublished data; D.A. Peltzer and S.D. bour mass only when grown in native vegetation,
Wilson, unpublished data). We examined grass and the growth of transplants of Agropyron
transplants with herbaceous neighbours, in order to decreased significantly with neighbour mass only
avoid drastic differences between transplant and when grown in Agropyron-dominated vegetation.
neighbour morphology (e.g. grasses vs. trees). We These results suggest that competition was most
excluded two young sites from Reader et al. (1994) intense in intraspecific pairings and that variation
because the topsoil had been removed and the plots in competitive effects occurred without variation in
were undergoing primary succession (H. Olff, abiotic stress. Similar results were found for the
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 201

Competition and Environmental Stress 201

Wilson and Tilman, 1991


Wilson and Tilman, 1993
2
Competition intensity

Wilson, 1993a
Reader et al., 1994
Wilson, 1994
1 2070
Gerry and Wilson, 1995
Wilson and Tilman, 1995
× Bakker, 1996
1205 Foster and Gross, 1997
0
× Peltzer et al., 1998
S.D. Wilson, unpublished data
× D.A. Peltzer and S.D. Wilson,
–1 unpublished data
0 200 400 600 700
Standing crop (g m–2)
Fig. 10.1. Competition intensity as a function of standing crop. Competition intensity was measured as the
relative reduction in grass transplant growth caused by grass neighbours. Standing crop increases with
decreasing stress. Data were taken from several studies in natural temperate grasslands.

grass Anthoxanthum odoratum in California, where in fertilized plots with less light penetration (Wilson
A. odoratum is more suppressed by intraspecific and Tilman 1991, 1993, 1995). This did not occur,
neighbours on dry soils than by interspecific neigh- however, in a grassland with a smaller range of
bours on moister soils (Platenkamp and Foin, standing crop (Belcher et al., 1995): in this case,
1990). Goldberg et al. (1995) outline a method for competition was always below ground, regardless of
testing whether differences in competition intensity neighbour mass. Competition was also entirely below
are attributable to differences in stress or simply to ground in species-poor experimental plots, in which
differences among neighbouring species. fertilization did not produce changes in species
composition (Peltzer et al., 1998). Taken together,
these studies suggest that light competition may
Stress and Root Competition occur only if standing crop is relatively high and if tall
life-forms are available to colonize fertilized plots and
If competition is equally important at all levels of cause shading.
stress, the mechanism of competition may still shift
from below ground to above ground as stress
decreases and standing crop increases. Field
experiments suggest that competition in grasslands Competitive Ability and Grassland
occurs mostly below ground and that neighbour Community Structure
shoots have little effect on transplant performance
(Cook, 1985; Snaydon and Howe, 1986; Wilson Variation in competitive ability among species
and Tilman, 1991, 1993, 1995; Seager et al., 1992; occurs at the individual level. A next step is to
Wilson, 1993a, b; Belcher et al., 1995; Peltzer et determine if variation among individual species can
al., 1998; see review by Casper and Jackson, 1997). be used to predict their performance in a community
Cook and Ratcliff (1984) used root exclusion tubes (Goldberg, 1990; Goldberg et al., 1995). Specifically,
of varying depths to show that the intensity of root does variation in the performance of individuals scale
competition decreased as neighbour-free soil vol- up to the population and community levels? One
ume increased. method to assess individual- vs. community-level
Shoot competition occurs in some grasslands if competitive ability is the community density series
standing crop is increased through fertilization. of Goldberg et al. (1995). This method manipulates
Transplant growth in unfertilized Minnesota old-field the density of the entire community to levels both
plots was controlled entirely by root competition but below and above those naturally occurring at a site.
was influenced by both root and shoot competition Very low-density treatments are null communities,
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 202

202 D.A. Peltzer and S.D. Wilson

which should have no competition occurring Interactions between Grazing


among plants. As the community density increases,
so do the frequency and importance of species and Competition
interactions, including competition. The difference
in performance of species between low- and high- Grazers have a number of impacts on native grass-
density treatments is its community-level competitive lands: they can remove 9–57% of net above-ground
ability. For example, species whose relative abundance foliage production (Frank et al., 1998), promote
increases with community density would have a shoot growth by removing older, less productive tis-
higher community-level competitive ability. Other sue (Caldwell et al., 1981; McNaughton, 1983,
general attempts to scale plant effects at lower scales 1984), enhance nutrient availability by increasing
to phenomena at higher scales have been discussed at rates of nutrient cycling (McNaughton et al., 1989;
length elsewhere (Ehleringer and Field, 1993; Jones Day and Detling, 1990; Holland et al., 1992) and
and Lawton, 1995; Bazzaz, 1996). create small-scale disturbances by trampling or bur-
Plant traits may be used to predict which species rowing (Crawley, 1983; Huntley, 1991; Olff and
have good competitive abilities, and to test these Ritchie, 1998). In addition, grazers have many
predictions empirically (Gaudet and Keddy, 1988; indirect effects in grasslands. For example, removal
Keddy, 1989, 1992; van der Werf et al., 1993; of above-ground biomass may enhance primary
Grime et al., 1997; Rösch et al., 1997; Reader, production through removal of detritus, increased
1998). A partial list of these traits is given in Table soil moisture status and plant water-use efficiency,
10.1. Whole communities can be screened for reduced fire frequency and altered species composi-
general patterns of traits associated with abundance tion (McNaughton, 1984, 1985; Knapp and
and distribution of species, although this is logisti- Seastedt, 1986; Archer, 1995).
cally difficult. Experimental evaluations of traits for At the level of the individual plant, grazers
competitive ability are needed (e.g. Wilson, 1991; reduce survival, growth and fecundity (Crawley,
Wedin and Tilman, 1993). 1983; Bullock, 1996). Several authors suggest that
One difficulty with trying to scale up variation grazing can promote plant growth through removing
in competitive ability or with using plant traits to senescing tissue and enhancing subsequent growth,
predict competitive ability is that competitive although the majority of studies do not support
ability may be inconsistent among sites or habitats this idea (see review by Belsky, 1986). By damaging
(Underwood and Petraitis, 1993; Miller, 1994; plants, herbivores induce defence mechanisms in
D.A. Peltzer and S.D. Wilson, unpublished data; plants such as altered morphology or increased
see review by Goldberg, 1996). For example, levels of defence chemicals in leaves (Crawley,
Vinton and Burke (1997) examined the effects of 1983; Vicari and Bazely, 1993). Grazing of above-
prairie plants on carbon and N cycling at three sites ground tissues also has consequences below ground.
varying in annual precipitation. They found that For example, classic work by Weaver (1950)
differences among species were greatest at the two showed that intense grazing of above-ground
most productive sites. This result was probably due foliage can strongly reduce below-ground biomass
to patchy plant cover at the driest site (the compari- in short-grass prairie. However, the effects of above-
son was between bare ground and plant cover) and ground grazing on below-ground productivity vary
a more continuous cover at the two wetter sites (the among grazing systems and the grazing history of the
comparison was among species). In a review of site (Belsky, 1986; Milchunas and Lauenroth, 1993).
studies comparing species’ competitive abilities Plants have several anti-herbivore defences
between environments, Goldberg (1996) found including high concentrations of silica in leaves,
that competitive responses tended to be consistent lower-growing meristems, high levels of fibre and
among environments while competitive effects low levels of protein and a variety of defence
often varied among environments. Further work is compounds (Vicari and Bazely, 1993). Several plant
needed to examine how species’ competitive traits confer resistance to both herbivory and
effects vary with environment and why. Factorial environmental stress. For example, high tissue
experiments designed to observe species effects  density, tough, fibrous leaves and high concentra-
environment interactions are ideal for determining tions of secondary compounds confer resistance to
variation in competitive abilities among species herbivores and stress (Chapin, 1980, 1991; Chapin
(Goldberg and Scheiner, 1993). et al., 1990; Grubb, 1992; see Table 10.1). In
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 203

Competition and Environmental Stress 203

addition, many traits allowing grasses to withstand unpalatable species assumes that a trade-off should
grazing may actually be adaptations to a semi-arid exist between competitive ability and grazing
environment (Coughenour, 1985). tolerance (Moretto and Distel, 1997). For example,
Grazing may have bigger impacts in stressful plant height should increase a plant’s competitive
habitats than in more productive habitats, for at ability for light, but may also increase its proneness
least two reasons. First, fewer nutrients are usually to grazing (Gaudet and Keddy, 1988; Oksanen,
available for plants to take up and use for regrowth 1993). A trade-off between competitive ability and
after biomass removal (Grime, 1979). Secondly, grazing tolerance should occur for competitive
shoot herbivory usually results in a decline in effect ability, but not necessarily for competitive
critical mycorrhizal mutualists associated with plant response ability; this is because many of the plant
roots (Gehring and Whitham, 1994), further traits conferring competition response or stress
reducing the ability of grazed plants to take up tolerance ability also confer resistance to herbivores
nutrients needed for compensatory growth. (see Table 10.1).
Herbivory may alter the outcome of competition Grazers may be essential for the restoration of
between grassland plant species. More specifically, native grasslands. Fire is often used in prairie
grazers can influence plant competitive ability restoration across North America, but repeated
either by modifying growth and morphology or by burning can lower plant species diversity (Collins et
influencing plant abundance and distribution al., 1995). Recent work by Collins et al. (1998)
(Louda et al., 1990). Further, by removing standing found that both mowing and grazing by bison
biomass, grazing acts to shift plant competition enhanced species diversity after fire by selectively
from shoots (i.e. for light) to roots (i.e. for soil reducing the abundance of dominant C4 (warm-
nutrients or water) (Milchunas et al., 1992). A season) grasses, thus reducing their competitive
common assertion is that herbivores shift the balance effects and allowing subordinate C3 (cool-season)
of competition by preferentially consuming domi- grasses and forbs to persist and contribute to grass-
nant competitors, altering the relative competitive land diversity. Restoration of diversity by grazers
ability of species and promoting the abundance of may be successful at small spatial scales (i.e.
plants that are otherwise rare or excluded (Tansley < 10–100 m2), but recent studies suggest that graz-
and Adamson, 1925; McNaughton, 1985; Huntley ing reduces plant diversity over larger spatial scales
1991; Clay et al., 1993; Olff and Ritchie, 1998). (i.e. > 1000 m2) (Olff and Ritchie, 1998; Stohlgren
This view is supported by a number of studies. For et al., 1999a, b).
example, selective grazing of the late-successional As discussed above, we know that grazing or
grass S. scoparium results in the increase of two herbivory strongly affects the survival, growth and
grasses that are normally poorer competitors, reproduction of plants, but how does herbivory
Bothriochloa saccharoides and Stipa leucotricha modify the outcome of competition along gradients
(Anderson and Briske, 1995). Grazing in a south- of stress (Connell, 1975; Menge and Sutherland,
ern English pasture increased the abundance of the 1987; Louda, 1989; Louda et al., 1990)? Most
N-fixing forb T. repens and decreased L. perenne, studies have examined the effects of grazers on
because the taller Lolium lost more above-ground range condition or species composition, or have
mass to grazers (Parsons et al., 1991). For two used manual clipping in pot experiments to
grasses growing in low-nutrient soils, clipping both examine the effects of grazers on plant competition;
caused greater reductions in the growth of A. elatius relatively few studies have quantified the effects of
than in that of Festuca rubra and accelerated the herbivory on plant competition under natural
replacement of Arrhenatherum by Festuca (Berendse conditions. Further, the effects of different kinds of
et al., 1992). Below-ground insect herbivory may herbivores on plant interactions is essentially
also accelerate succession to a grass-dominated unknown, i.e. what are the effects of large-bodied
sward by reducing the size and growth of perennial ungulates vs. small burrowing mammals vs. insects,
forbs and perhaps their competitive ability (Brown or native grazers vs. cattle in the same grassland
and Gange, 1990, 1992). Thus, herbivore-induced system? For example, even though below-ground
changes of competitive interactions can drive herbivory can greatly reduce plant productivity, few
species replacement in grasslands. studies have considered the role of invertebrate
The argument that selective grazing causes the grazers on roots, even though most grassland
competitive replacement of palatable species by production is below ground (Stanton, 1988; Brown
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 204

204 D.A. Peltzer and S.D. Wilson

and Gange, 1990). In short-grass prairie, herbivory responded similarly to water stress. The authors
by white grubs, the larvae of June beetles (including suggest that niche separation may occur along
Phyllophaga fimbripes), strongly depresses the abun- seasonal temperature gradients rather than seasonal
dance of the dominant grass, B. gracilis, for up to moisture gradients for these two grasses. Sala et al.
14 years after an insect outbreak (Coffin et al., (1982) found that A. smithii took up soil moisture
1998). Clearly, further work is required to under- earlier in the growing season than did B. gracilis.
stand how different kinds of above- and below- Similar results were reported for the same two
ground herbivory interact with competition to grasses by Christie and Detling (1982), using a
determine the productivity and species composition replacement series experiment. Similarly, many
of grasslands. authors suggest that the success of the invasive
grass Bromus tectorum in large areas of the western
USA is at least partly due to its growth earlier in
Non-competitive Interactions the season than native grasses and shrubs (Mack,
1986). Niche space can also be divided up either
Many plant species can coexist in grasslands despite through different regeneration niches (Grubb,
having large differences in competitive abilities (e.g. 1977) or through ratios of limiting nutrients
Mitchley and Grubb, 1986; Silvertown et al., (Tilman, 1982).
1994), suggesting that other factors either prevent Because there is no theory predicting which
the competitive exclusion of species or promote niche axes species should segregate along, it is
their coexistence (Bullock, 1996). impossible to determine if no niche segregation
Grasses might avoid competition by exploiting occurs among species or whether the wrong niche
different niches. If this happens, then competition axes were measured. Thus, there are difficulties with
might contribute little to grassland patterns interpreting what niche differences mean when
(Leibold, 1995). It is difficult to imagine how they are found and whether they are the cause or
plants can divide up resources into distinct niches, the result of species coexistence (Arthur, 1982;
because all plants use the same resources (Harper, Silvertown, 1983). Leibold (1995) linked tradi-
1977; Tilman, 1982; Goldberg and Werner, 1983). tional theory of the niche with mechanistic models
Several studies have explicitly examined resource of community interactions (e.g. Tilman, 1982;
partitioning in plant communities (e.g. Werner and Holt et al., 1994), which may be a more promising
Platt, 1976; Russell et al., 1985), but are often approach to testing niche theory.
inconclusive. For example, Mahdi et al. (1988) did More recent alternative explanations of species
not find significant niche separation among eight coexistence have been developed which do not
British grassland species using either six niche invoke niche differentiation. Silvertown and Law
dimensions based on soil nutrient levels and (1987) review some of these, including spatial
phenology or ratios of N : P as a test of Tilman’s aggregation in populations (Yodzis, 1986), non-
(1982) resource ratio hypothesis. transitive competition (Keddy, 1989; Shipley,
In North American prairies, many species 1993), a variety of non-equilibrium theories
coexist in relatively few habitat types. In these (Shmida and Ellner, 1984; Chesson and Huntly,
grasslands, two niche axes are commonly studied: 1997) and density-independent mortality (Huston,
water use and phenology. Werner and Platt (1976) 1979). For example, Ågren and Fagerström (1984)
showed niche partitioning for populations of six demonstrated that two species with similar juvenile
goldenrod species (Solidago spp.) along a soil mois- competitive abilities and lifespan can coexist if
ture gradient in an Iowa prairie, although the same seed production (i.e. reproductive events) varies
species did not show niche partitioning in an old stochastically. Other models have demonstrated
field in Michigan. Many grassland species display that spatial patchiness or environmental hetero-
phenological divergence in growth, which is related geneity promotes species coexistence, at least for
to photosynthetic pathway (Ehleringer and annual communities (Pacala and Tilman, 1994). In
Monson, 1993). For example, Kemp and Williams a recent review, Chesson (1991) suggests that we
(1980) examined photosynthesis, respiration and look for trade-offs that can determine coexistence
growth in two common prairie grasses, Agropyron and stability in communities – for example,
smithii (C3) and B. gracilis (C4). Bouteloua gracilis dispersal vs. competitive ability in grassland plants
grew best in warmer conditions, but both species (Tilman, 1997).
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 205

Competition and Environmental Stress 205

Conclusions linked to the abundance or distribution of species.


In contrast, grasses that are best able to reduce
Field experiments suggest that competition controls resources (i.e. good effect competitors) tend to
plant performance at all levels of stress in natural dominate nutrient-poor soils, suggesting that com-
grasslands (see Fig. 10.1). Further, competition is petitive effect ability may determine the relative
primarily among roots for below-ground resources, abundance of species in grasslands.
such as water and nutrients. The relative importance In the process of resource consumption, grasses
of shoot competition can increase as stress decreases, may also affect the patchiness of resources. The
shoot mass increases and light penetration decreases. scale of this patchiness is in turn likely to influence
In addition, results from previous studies indicate other plants. Our understanding of the relative
that competition intensity changes more along contributions of competition and stress to pasture
natural stress gradients than along experimental community composition will be enhanced by closer
gradients: we need studies of the differences study of the dynamics and heterogeneity of below-
between natural and experimental stress gradients ground processes.
in the same system. Particularly for natural gradi- Understanding how competitive ability and
ents, we need to separate the effects of abiotic and environment interact is increasingly important
biotic stress, in order to determine the relative because grassland environments are often altered by
role of environment and species interactions in humans through fertilization, nutrient pollution,
controlling the species composition in grassland mowing, grazing and the introduction of new
vegetation. species. Moderate levels of grazing can interact
Using plant traits to predict individual- and with plant competition to maintain the diversity
community-level competitive ability is a powerful of grasslands and offset declines in plant diversity
approach to testing current competition theory. caused by frequent fires. Thus, grazing may be an
Screening experiments using several species in important tool for the restoration of grassland
environments differing in stress or productivity are diversity. However, more work is needed on
needed; interactions among species, competition comparing the effects of different grazers on the
and environment should be examined to determine outcome of competition in the same system.
how competitive ability changes with abiotic and
biotic stresses. Results to date indicate that many of
the traits that characterize stress-tolerant species Acknowledgements
should also enable such species to be effective
competitors in stressful habitats. We thank the Natural Sciences and Engineering
Field experiments have failed to show that an Research Council of Canada for support and
ability to perform well in the presence of neigh- P. Tow and A. Lazenby for helpful comments on an
bours (i.e. good competitive response ability) is earlier draft of the manuscript.

References
Aarssen, L.W. (1988) ‘Pecking order’ of four plant species from pastures of different ages. Oikos 51, 3–12.
Aarssen, L.W. (1992) Causes and consequences of variation in competitive ability in plant communities. Journal of
Vegetation Science 3, 165–174.
Aerts, R. and van der Peijl, M.J. (1993) A simple model to explain the dominance of low-productive perennials in nutrient-
poor environments. Oikos 66, 144–147.
Ågren, G.I. and Fagerström, T. (1984) Limiting dissimilarity in plants: randomness prevents exclusion of species with
similar competitive abilities. Oikos 43, 369–375.
Allen, E.B. and Allen, M.F. (1990) The mediation of competition by mycorrhizae in successional and patchy environments.
In: Grace, J.B. and Tilman, D. (eds) Perspectives on Plant Competition. Academic Press, San Diego, USA,
pp. 367–389.
Anderson, V.J. and Briske, D.D. (1995) Herbivore-induced species replacement in grasslands: is it driven by herbivory
tolerance or avoidance? Ecological Applications 5, 1014–1024.
Archer, S. (1995) Tree-grass dynamics in a Prosopis-thornscrub savanna parkland: reconstructing the past and predicting
the future. Ecoscience 2, 83–99.
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 206

206 D.A. Peltzer and S.D. Wilson

Arthur, W. (1982) The evolutionary consequences of interspecific competition. In: MacFayden, A. and Ford, E.D. (eds)
Advances in Ecological Research. Academic Press, London, UK, pp. 127–187.
Austin, M.P. (1990 ) Community theory and competition in vegetation. In: Grace, J.B. and Tilman, D. (eds) Perspectives
on Plant Competition. Academic Press, New York, USA, pp. 215–238.
Bakker, J.D. (1996) Competition and the establishment of native grasses in crested wheatgrass fields. MSc thesis,
University of Regina, Saskatchewan, Canada.
Bazzaz, F.A. (1996) Plants in Changing Environments: Linking Physiological, Population, and community Ecology.
Cambridge University Press, New York, USA.
Belcher, J.W., Keddy, P.A. and Twolan-Strutt, L. (1995) Root and shoot competition intensity along a soil depth gradient.
Journal of Ecology 83, 673–682.
Bell, G. and Lechowicz, M.J. (1994) Spatial heterogeneity at small scales and how plants respond to it. In: Caldwell,
M.M. and Pearcy, R.W. (eds) Exploitation of Environmental Heterogeneity by Plants. Academic Press, San Diego,
USA, pp. 391–414.
Belsky, A.J. (1986) Does herbivory benefit plants? A review of the evidence. American Naturalist 127, 870–892.
Berendse, F. (1994) Competition between plant populations at low and high nutrient supplies. Oikos 71, 253–260.
Berendse, F. and Aerts, R. (1987) Nitrogen-use efficiency: a biologically meaningful definition. Functional Ecology 1,
293–296.
Berendse, F. and Elberse, W. (1990) Competition and nutrient availability in the heathland and grassland ecosystems.
In: Grace, J.B. and Tilman, D. (eds) Perspectives in Plant Competition. Academic Press, San Diego, USA,
pp. 93–116.
Berendse, F., Elberse, W.Th. and Geerts, R.H.M.E. (1992) Competition and nitrogen loss from plants in grassland
ecosystems. Ecology 73, 46–53.
Bever, J.D. (1994) Feedback between plants and their soil communities in an old field community. Ecology 75,
1965–1977.
Bever, J.D., Westover, K.M. and Antonovics, J. (1997) Incorporating the soil community into plant population dynamics:
the utility of the feedback approach. Journal of Ecology 85, 561–573.
Bilbrough, C.J. and Caldwell, M.M. (1997) Exploitation of springtime ephemeral N pulses by six Great Basin plant
species. Ecology 78, 231–243.
Bonser, S.P. and Reader, R.J. (1995) Plant competition and herbivory in relation to vegetation biomass. Ecology 76,
2176–2183.
Brown, V.K. and Gange, A.C. (1990) Insect herbivory belowground. Advances in Ecological Research 20, 1–58.
Brown, V.K. and Gange, A.C. (1992) Secondary plant succession: how is it modified by insect herbivory? Vegetatio 101,
3–13.
Bullock, J.M. (1996) Plant competition and population dynamics. In: Hodgson, J. and Illius, A.W. (eds) The Ecology
and Management of Grazing Systems. CAB International, Wallingford, UK, pp. 69–100.
Caldwell, M.M. (1994) Exploiting nutrients in fertile soil microsites. In: Caldwell, M.M. and Pearcy, R.W. (eds)
Exploitation of Environmental Heterogeneity by Plants. Academic Press, San Diego, USA, pp. 325–347.
Caldwell, M.M. and Pearcy, R.W. (1994) Exploitation of Environmental Heterogeneity by Plants: Ecophysiological Processes
Above and Below Ground. Academic Press, New York, USA.
Caldwell, M.M. and Richards, J.H. (1986) Competing root systems: morphology and models of absorption. In:
Givnish, T.J. (ed.) On the Economy of Plant Form and Function. Cambridge University Press, Cambridge, UK,
pp. 251–273.
Caldwell, M.M., Richards, J.H., Johnson, D.A., Nowak, R.S. and Dzuree, R.S. (1981) Coping with herbivory: photo-
synthetic capacity and resource allocation in two semiarid Agropyron bunchgrasses. Oecologia 50, 14–24.
Campbell, B.D. and Grime, J.P. (1989) A comparative study of plant responsiveness to the duration of episodes of mineral
nutrient enrichment. New Phytologist 112, 261–267.
Campbell, B.D., Grime, J.P. and Mackey, J.M.L. (1991) A tradeoff between scale and precision in resource foraging.
Oecologia 87, 532–538.
Casper, B.B. and Cahill, J.F. (1998) Population level responses to nutrient heterogeneity and density by Abutilon
theophrasti (Malvaceae): an experimental neighborhood approach. American Journal of Botany 85, 1680–1687.
Casper, B.B. and Jackson, R.B. (1997) Plant competition underground. Annual Review of Ecology and Systematics 28,
545–570.
Chapin, F.S., III (1980) The mineral nutrition of wild plants. Annual Review of Ecology and Systematics 11, 233–260.
Chapin, F.S., III (1991) Effects of multiple environmental stresses on nutrient availablity and use. In: Mooney, H.A.,
Winner, W.E. and Pell, E.J. (eds) Response of Plants to Multiple Stresses. Academic Press, San Diego, USA,
pp. 67–88.
Chapin, F.S., III, Schultze, E.-D. and Mooney, H.A. (1990) The ecology and economics of storage in plants. Annual
Review of Ecology and Systematics 21, 423–447.
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 207

Competition and Environmental Stress 207

Chapin, F.S., III, Autumn, K. and Pugnaire, F. (1993) Evolution of suites of traits in response to environmental stress.
American Naturalist 142, 78–92.
Cheplick, G.P. (1997) Responses to severe competitive stress in a clonal plant: differences between genotypes. Oikos 79,
581–591.
Chesson, P. (1991) A need for niches? Trends in Ecology and Evolution 6, 26–28.
Chesson, P. and Huntly, N. (1997) The roles of harsh and fluctuating conditions in the dynamics of ecological communities.
American Naturalist 150, 519–553.
Christie, E.K. and Detling, J.K. (1982) Analysis of interference between C3 and C4 grasses in relation to temperature
and soil nitrogen supply. Ecology 63, 1277–1284.
Clay, K., Marks, S. and Cheplick, G.P. (1993) Effects of insect herbivory and fungal endophyte infection on competitive
interactions among grasses. Ecology 74, 1767–1777.
Coffin, D.P., Laycock, W.A. and Lauenroth, W.K. (1998) Disturbance intensity and above- and belowground herbivory
effects on long-term (14 y) recovery of a semiarid grassland. Plant Ecology 139, 221–233.
Collins, S.L., Glenn, S.M. and Gibson, D.J. (1995) Experimental analysis of intermediate disturbance and initial floristic
composition: decoupling cause and effect. Ecology 76, 486–492.
Collins, S.L., Knapp, A.K., Briggs, J.M., Blair, J.M. and Steinauer, E.M. (1998) Modulation of diversity by grazing and
mowing in native tallgrass prairie. Science 280, 745–747.
Connell, J.H. (1975) Some mechanisms producing structure in natural communities: a model and evidence from field
experiments. In: Cody, M.L. and Diamond, J.M. (eds) Ecology and Evolution of Communities. Belknap Press,
Cambridge, UK, pp. 460–490.
Cook, S.J. (1985) Effect of nutrient application and herbicides on root competition between green panic seedlings and a
Heteropogon grassland sward. Grass and Forage Science 40, 171–175.
Cook, S.J. and Ratcliff, D. (1984) A study of the effects of root and shoot competition on the growth of green panic
(Panicum maximum var. trichoglume) seedlings in an existing grassland using root exclusion tubes. Journal of
Applied Ecology 21, 971–982.
Coughenour, M.B. (1985) Graminoid responses to grazing by large herbivores: adaptations, exaptation, and interacting
processes. Annals of the Missouri Botanical Garden 72, 852–863.
Crawley, M.J. (1983) Herbivory: the dynamics of animal-plant interations. Studies in Ecology Vol. 10, University of
California Press, Berkeley, California, USA.
Day, T.A. and Detling, J.K. (1990) Grassland patch dynamics and herbivore grazing preference following urine deposition.
Ecology 71, 180–188.
Del Moral, R. (1983) Competition as a control mechanism in subalpine meadows. American Journal of Botany 70,
232–245.
Diamond, J.M. (1975) Assembly of species communities. In: Cody, M.L. and Diamond, J.M. (eds) Ecology and
Evolution of Communities. Belknap Press, Cambridge, UK, pp. 342–444.
DiTommaso, A. and Aarssen, L.W. (1991) Effect of nutrient level on competition intensity in the field for three coexisting
grass species. Journal of Vegetation Science 2, 513–522.
Ehleringer, J.R. and Field, C.B. (1993) Scaling Physiological Processes: Leaf to Globe. Academic Press, Orlando, USA.
Ehleringer, J.R. and Monson, R.K. (1993) Evolutionary and ecological aspects of photosynthetic pathway variation.
Annual Review of Ecology and Systematics 24, 411–439.
Felsenstein, J. (1988) Phylogenies and quatitative methods. Annual Review of Ecology and Systematics 19, 445–471.
Fitter, A.H. (1986) Aquisition and utilization of resources. In: Crawley, M.J. (ed.) Plant Ecology. Blackwell, Oxford, UK,
pp. 375–405.
Fitter, A.H. (1991) Costs and benefits of mycorrhizas: implications for functioning under natural conditions.
Experientia 47, 350–355.
Foster, B.L. and Gross, K.L. (1997) Partitioning the effects of plant biomass and litter on Andropogon gerardii in old-
field vegetation. Ecology 2091–2104,
Frank, D.A., McNaughton, S.J. and Tracy, B.F. (1998) The ecology of the earth’s grazing ecosystems. BioScience 48,
513–521.
Gaudet, C.L. and Keddy, P.A. (1988) A comparative approach to predicting competitive ability from plant traits. Nature
334, 242–243.
Gehring, C.A. and Whitham, T.G. (1994) Interactions between aboveground herbivores and the mycorrhizal mutualists
of plants. Trends in Ecology and Evolution 9, 251–255.
Gerry, A.K. and Wilson, S.D. (1995) The influence of initial size on the competitive responses of six plant species.
Ecology 76, 272–279.
Givnish, T.J. (1986) Economics of biotic interactions. In: Givnish, T.J. (ed.) On the Economy of Plant Form and
Function. Cambridge University Press, Cambridge, UK, pp. 667–680.
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 208

208 D.A. Peltzer and S.D. Wilson

Goldberg, D.E. (1990) Components of resource competition in plant communities. In: Grace, J.B. and Tilman, D.
(eds) Perspectives on Plant Competition. Academic Press, San Diego, USA, pp. 27–49.
Goldberg, D.E. (1996) Competitive ability: definitions, contingency and correlated traits. Philosophical Proceedings of the
Royal Society of London B 351, 1377–1385.
Goldberg, D.E. and Barton, A.M. (1992) Patterns and consequences of interspecific competition in natural communities:
a review of field experiments with plants. American Naturalist 139, 771–801.
Goldberg, D. and Novoplansky, A. (1997) On the relative importance of competition in unproductive environments.
Journal of Ecology 85, 409–418.
Goldberg, D.E. and Scheiner, S.M. (1993) ANOVA and ANCOVA: field competition experiments. In: Scheiner, S.M.
and Gurevitch, J. (eds) Design and Analysis of Ecological Experiments. Chapman and Hall, New York, USA,
pp. 69–93.
Goldberg, D.E. and Werner, P.A. (1983) Equivalence of competitors in plant communities: a null hypothesis and a field
experimental approach. American Journal of Botany 70, 1098–1104.
Goldberg, D.E., Turkington, R. and Olsvig-Whittaker, L. (1995) Quantifying the community-level consequences of
competition. Folia Geobotanica et Phytotaxonomica 30, 231–242.
Grace, J.B. (1990) On the relationship between plant traits and competitive ability. In: Grace, J.B. and Tilman, D. (eds)
Perspectives on Plant Competition. Academic Press, San Diego, USA, pp. 51–65.
Grace, J.B. (1991) A clarification of the debate between Grime and Tilman. Functional Ecology 5, 583–587.
Grace, J.B. (1995) On the measurement of plant competition intensity. Ecology 76, 305–307.
Greenlee, J.T. and Callaway, R.M. (1996) Abiotic stress and the importance of interference and facilitation in montane
bunchgrass communities in western Montana. American Naturalist 148, 386–396.
Grime, J.P. (1977) Evidence for the existence of three primary strategies in plants and its relevance to ecological and
evolutionary theory. American Naturalist 111, 1169–1194.
Grime, J.P. (1979) Plant Strategies and Vegetation Processes. John Wiley & Sons, Chichester, UK.
Grime, J.P. (1994) The role of plasticity in exploiting environmental heterogeneity. In: Caldwell, M.M. and Pearcy,
R.W. (eds) Exploitation of Environmental Heterogeneity by Plants. Academic Press, San Diego, USA, pp. 1–19.
Grime, J.P., Thompson, K., Hunt, R., Hodgson, J.G., Cornelissen, J.H.C., Rorison, I.H., Hendry, G.A.F., Ashenden,
T.W., Askew, A.P., Band, S.R., Booth, R.E., Bossard, C.C., Campbell, B.D., Cooper, J.E.L., Davison, A.W.,
Gupta, P.L., Hall, W., Hand, D.W., Hannah, M.A., Hillier, S.H., Hodkinson, D.J., Jalili, A., Liu, Z., Mackey,
J.M.L., Matthews, N., Mowforth, M.A., Neal, A.M., Reader, R.J., Reiling, K., Ross-Fraser, W., Spencer, R.E.,
Sutton, F., Tasker, D.E., Thorpe, P.C. and Whitehouse, J. (1997) Integrated screening validates primary axes of
specialisation in plants. Oikos 79, 259–281.
Grubb, P.J. (1977) The maintenance of species-richness in plant communities: the importance of the regeneration niche.
Biological Reviews 52, 107–145.
Grubb, P.J. (1992) A positive distrust in simplicity – lessons from plant defences and from competition among plants
and among animals. Journal of Ecology 80, 585–610.
Gurevitch, J. (1986) Competition and the local distribution of the grass Stipa neomexicana. Ecology 67, 46–57.
Gurevitch, J., Morrow, L.A., Wallace, A. and Walsh, J.S. (1992) A meta-analysis of competition in field experiments.
American Naturalist 140, 539–572.
Harper, J.L. (1965) The nature and consequences of the interference amongst plants. Proceedings of the International
Conference on Genetics 11, 465–481.
Harper, J.L. (1977) Population Biology of Plants. Academic Press, London, UK.
Harvey, P.H. and Pagel, M.D. (1991) The comparative Method in Evolutionary Biology. Oxford University Press,
Oxford, UK.
Herben, T. and Krahulec, F. (1990) Competitive hierarchies, reversals of rank order and the de Wit approach: are they
compatible? Oikos 58, 254–256.
Holland, E.A., Parton, W.J., Detling, J.K. and Coppock, D.L. (1992) Physiological responses of plant populations to
herbivory and their consequences for ecosystem nutrient flow. American Naturalist 140, 685–706.
Holt, R.D., Grover, H. and Tilman, D. (1994) Simple rules for interspecific dominance in systems with exploitative and
apparent competition. American Naturalist 144, 741–771.
Hook, P.B., Burke, I.C. and Lauenroth, W.K. (1991) Heterogeneity of soil and plant N and C associated with individual
plants and openings in North American shortgrass steppe. Plant and Soil 138, 247–256.
Humphrey, L.D. and Pyke, D.A. (1997) Clonal foraging in perennial wheatgrasses: a strategy for exploiting patchy soil
nutrients. Journal of Ecology 85, 601–610.
Huntley, N.J. (1991) Herbivores and the dynamics of communities and ecosystems. Annual Review of Ecology and
Systematics 22, 477–503.
Huston, M. (1979) A general hypothesis of species diversity. American Naturalist 113, 81–101.
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 209

Competition and Environmental Stress 209

Huston, M. and Smith, T. (1987) Plant succession: life history and competition. American Naturalist 130, 168–198.
Huston, M.A. and DeAngelis, D.L. (1994) Competition and coexistence: the effects of resource transport and supply
rates. American Naturalist 144, 954–977.
Hutchings, M.J. (1988) Differential foraging for resources and structural plasticity in plants. Trends in Ecology and
Evolution 3, 200–204.
Jackson, L.E., Strauss, R.B., Firestone, M.K. and Bartolome, J.W. (1990) Influence of tree canopies on grassland pro-
ductivity and nitrogen dynamics in deciduous oak savanna. Agricultural Ecosystems and Environments 32, 89–105.
Jones, C. and Lawton, J.H. (1995) Linking Species and Ecosystems. Chapman and Hall, New York, USA.
Kadmon, R. (1995) Plant competition along soil moisture gradients: a field experiment with the desert annual Stipa
capensis. Journal of Ecology 83, 253–262.
Keddy, P.A. (1989) Competition. Chapman and Hall, London, UK.
Keddy, P.A. (1992) Assembly and response rules: two goals for predictive community ecology. Journal of Vegetation
Science 3, 157–164.
Keddy, P.A. and Shipley, B. (1989) Competitive hierarchies in herbaceous plant communities. Oikos 54, 234–241.
Kemp, P.R. and Williams, G.J. (1980) A physiological basis for niche separation between Agropyron smithii (C3) and
Bouteloua gracilis (C4). Ecology 61, 846–858.
Kleb, H.R. and Wilson, S.D. (1997) Vegetation effects on soil resource heterogeneity in prairie and forest. American
Naturalist 150, 283–298.
Knapp, A.K. and Seastedt, T.R. (1986) Detritus accumulation limits productivity of tallgrass prairie. BioScience 36,
662–668.
Lauenroth, W.K. and Aguilera, M.O. (1998) Plant-plant interactions in grasses and grasslands. In: Cheplick, G.P. (ed.)
Population Biology of Grasses. Cambridge University Press, Cambridge, UK, pp. 209–230.
Lauenroth, W.K., Dodd, J.L. and Sims, P.L. (1978) The effects of water- and nitrogen-induced stresses on plant
community structure in a semiarid grassland. Oecologia 36, 211–222.
Leibold, M.A. (1995) The niche concept revisited: mechanistic models and community coexistence. Ecology 76,
1371–1382.
Li, X. and Wilson, S.D. (1998) Facilitation among woody plants establishing in an old field. Ecology 79, 2694–2705.
Louda, S. (1989) Differential predation pressure: a general mechanism for structuring plant communities along complex
gradients? Trends in Ecology and Evolution 4, 158–159.
Louda, S.M., Keeler, K.H. and Holt, R.D. (1990) Herbivore influences on plant performance and competitive interac-
tions. In: Grace, J. and Tilman, D. (eds) Perspectives in Plant Competition. Academic Press, San Diego, USA,
pp. 414–444.
Mack, R.N. (1986) Alien plant invasion into the intermountain west: a case history. In: Mooney, H.A. and Drake, J.
(eds) Ecology of Biological Invasions of North America and Hawaii. Springer-Verlag, New York, USA, pp. 191–213.
McLellan, A.J., Law, R. and Fitter, A.H. (1997) Response of calcareous grassland plant species to diffuse competition:
results from a removal experiment. Journal of Ecology 85, 479–490.
McNaughton, S.J. (1983) Compensatory plant growth as a response to herbivory. Oikos 40, 329–336.
McNaughton, S.J. (1984) Grazing lawns: animals in herds, plant form, and coevolution. American Naturalist 124,
863–886.
McNaughton, S.J. (1985) Ecology of a grazing ecosystem: the Serengeti. Ecological Monographs 55, 259–294.
McNaughton, S.J., Oesterheld, M., Frank, D.A. and Williams, K.J. (1989) Ecosystem-level patterns of primary produc-
tivity and herbivory in terrestrial habitats. Nature 341, 142–144.
Mahdi, A., Law, R. and Willis, A.J. (1988) Large niche overlaps among coexisting plant species in a limestone grassland
community. Journal of Ecology 76, 386–400.
Mehrhoff, L.A. and Turkington, R. (1990) Microevolution and site-specific outcomes of competition among pasture
plants. Journal of Ecology 78, 745–756.
Menge, B.A. and Sutherland, J.P. (1987) Community regulation: variation in disturbance, competition, and predation
in relation to environmental stress and recruitment. American Naturalist 130, 730–757.
Milchunas, D.G. and Lauenroth, W.K. (1993) Quantitative effects of grazing on vegetation and soils over a global range
of environments. Ecological Monographs 63, 327–366.
Milchunas, D.G., Lauenroth, W.K. and Chapman, P.L. (1992) Plant competition, abiotic, and long- and short-term
effects of large herbivores on demography of opportunistic species in a semiarid grassland. Oecologia 92, 520–531.
Miller, R.E., Ver Hoeff, J.M. and Fowler, N.L. (1995) Spatial heterogeneity in eight central Texas grasslands. Journal of
Ecology 83, 919–928.
Miller, T.E. (1994) Direct and indirect species interactions in an early old-field plant community. American Naturalist
143, 1007–1025.
Miller, T.E. (1996) On quantifying the intensity of competition across gradients. Ecology 77, 978–981.
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 210

210 D.A. Peltzer and S.D. Wilson

Mitchley, J. and Grubb, P.J. (1986) The control of relative abundance of perennials in chalk grassland in southern
England. I. Constancy of rank order and results of pot and field experiments on the role of interference. Journal of
Ecology 74, 1139–1166.
Moretto, A.S. and Distel, R.A. (1997) Competitive interactions between palatable and unpalatable grasses native to a
temperate semi-arid grassland of Argentina. Plant Ecology 130, 155–161.
Mouat, M.C.H. (1983) Competitive adaptation by plants to nutrient shortage through modification of root growth and
surface change. New Zealand Journal of Agricultural Science 26, 327–332.
Newman, E.I. (1973) Competition and diversity in herbaceous vegetation. Nature 244, 310.
Oksanen, L. (1993) Plant strategies and environmental stress: a dialectic approach. In: Fowden, L., Mansfield, T. and
Stoddart, J. (eds) Plant Adaptation to Environmental Stress. Chapman and Hall, London, UK, pp. 313–333.
Olff, H. and Ritchie, M.E. (1998) Effects of herbivores on grassland plant diversity. Trends in Ecology and Evolution 13,
261–265.
Pacala, S.W. and Tilman, D. (1994) Limiting similarity in mechanistic and spatial models of plant competition in
heterogeneous environments. American Naturalist 143, 222–257.
Parsons, A.J., Harvey, A. and Johnson, I.R. (1991) Plant–animal interactions in a continuously grazed mixture. 2. The
role of differences in the physiology of plant growth and of selective grazing on the performance and stability of
species in a mixture. Journal of Applied Ecology 28, 635–658.
Peltzer, D.A., Wilson, S.D. and Gerry, A.K. (1998) Competition intensity along a productivity gradient in a low-
diversity grassland. American Naturalist 151, 465–476.
Platenkamp, G.A.V. and Foin, T.C. (1990) Ecological and evolutionary importance of neighbors in the grass
Anthoxanthum odoratum. Oecologia 83, 201–208.
Read, D.J. (1991) Mycorrhizas in ecosystems. Experientia 47, 376–391.
Reader, R.J. (1990) Competition constrained by low nutrient supply: an example involving Hieracium floribundum
Wimm and Grab. (Compositae). Functional Ecology 4, 573–577.
Reader, R.J. (1992) Herbivory as a confounding factor in an experiment measuring competition among plants. Ecology
73, 373–376.
Reader, R.J. (1993) Control of seedling emergence by ground cover and seed predation in relation to seed size for some
old-field species. Journal of Ecology 81, 169–176.
Reader, R.J. (1998) Relationship between species relative abundance and plant traits for an infertile habitat. Plant
Ecology 134, 43–51.
Reader, R.J. and Best, B.J. (1989) Variation in competition along an environmental gradient: Hieracium floribundum in
an abandoned pasture. Journal of Ecology 77, 673–684.
Reader, R.J. and Bonser, S.P. (1993) Control of plant frequency on an environmental gradient – effects of abiotic
variables, neighbours, and predators on Poa pratensis and Poa compressa (Gramineae). Canadian Journal of Botany
71, 592–597.
Reader, R.J., Wilson, S.D., Belcher, J.W., Wisheu, I., Keddy, P.A., Tilman, D., Morris, E.C., Grace, J.B., McGraw, J.B.,
Olff, H., Turkington, R., Klein, E., Leung, Y., Shipley, B., van Hulst, R., Johansson, M.E., Nilsson, C., Gurevitch,
J., Grigulis, K. and Beisner, B.E. (1994) Intensity of plant competition in relation to neighbor biomass: an inter-
continental study with Poa pratensis. Ecology 75, 1753–1760.
Reynolds, H.L., Hungate, B.A., Chapin, F.S., III and D’Antonio, C.M. (1997) Soil heterogeneity and plant competition
in an annual grassland. Ecology 78, 2076–2090.
Rösch, H., Van Rooyen, M.W. and Theron, G.K. (1997) Predicting competitive interactions between pioneer plant
species by using plant traits. Journal of Vegetation Science 8, 489–499.
Russell, P.J., Flowers, T.J. and Hutchings, M.J. (1985) Comparison of niche breadths and overlaps of halophytes on salt
marshes of differing diversity. Vegetatio 61, 171–178.
Sala, O.E. and Lauenroth, W.K. (1985) Root profiles and the ecological effect of light rainshowers in arid and semiarid
regions. American Midland Naturalist 114, 406–408.
Sala, O.E., Lauenroth, W.K. and Reid, C.P.P. (1982) Water relations: a new dimension for niche separation between
Bouteloua gracilis and Agropyron smithii in North American semiarid grasslands. Journal of Applied Ecology 19,
647–657.
Sala, O.E., Parton, W.J., Joyce, L.A. and Lauenroth, W.K. (1988) Primary production of the central grassland region of
the United States. Ecology 69, 40–45.
Sala, O.E., Lauenroth, W.K. and Parton, W.J. (1992) Long-term soil water dynamics in the shortgrass steppe. Ecology
73, 1175–1181.
Seager, N.G., Kemp, P.D. and Chu, A.C.P. (1992) Effect of root and shoot competition from established hill country
pasture on perennial ryegrass. New Zealand Journal of Agricultural Research 35, 359–363.
Shipley, B. (1993) A null model for competitive hierarchies in competition matrices. Ecology 74, 1693–1699.
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 211

Competition and Environmental Stress 211

Shipley, B. and Keddy, P.A. (1993) Evaluating the evidence for competitive hierarchies in plant communities. Oikos 69,
340–345.
Shmida, A. and Ellner, S. (1984) Coexistence of plant species with similar niches. Vegetatio 58, 29–55.
Shorrocks, B. and Swingland, I.R. (1990) Living in a Patchy Environment. Oxford University Press, Oxford, UK.
Silvertown, J. (1983) The distribution of plant species in limestone pavement: tests of species interaction and niche sep-
aration against null hypotheses. Journal of Ecology 71, 819–828.
Silvertown, J. and Dale, P. (1991) Competitive hierarchies and the structure of herbaceous plant communities. Oikos 61,
441–444.
Silvertown, J. and Dodd, M. (1996) Comparing plants and connecting traits. Philisophical Transactions of the Royal
Society of London B 351, 1233–1239.
Silvertown, J. and Law, R. (1987) Do plants need niches? Some recent developments in plant community ecology.
Trends in Ecology and Evolution 2, 24–26.
Silvertown, J., Dodd, M.E., McConway, K., Potts, J. and Crawley, M. (1994) Rainfall, biomass variation, and
community composition in the Park Grass Experiment. Ecology 75, 2430–2437.
Smith, T.M. and Huston, M.A. (1989) A theory of the spatial and temporal dynamics of plant communities. Vegetatio
83, 49–69.
Snaydon, R.W. and Howe, D.C. (1986) Root and shoot competition between established ryegrass and invading grass
seedlings. Journal of Applied Ecology 23, 667–674.
Stanton, N.L. (1988) The underground in grasslands. Annual Review of Ecology and Systematics 19, 573–589.
Stark, J.M. (1994) Causes of soil nutrient heterogeneity at different scales. In: Caldwell, M.M. and Pearcy, R.W. (eds)
Exploitation of Environmental Heterogeneity by Plants. Academic Press, San Diego, USA, pp. 255–284.
Stohlgren, T.J., Binkley, D., Chong, G.W., Kalkhan, M.A., Schell, L.D., Bull, K.A., Otsuki, Y., Newman, G., Bashkin,
M. and Son, Y. (1999a) Exotic plant species invade hot spots of native plant diversity. Ecological Monographs 69,
25–46.
Stohlgren, T.J., Schell, L.D. and Vanden Heuvel, B. (1999b) How grazing and soil quality affect native and exotic plant
diversity in Rocky Mountain grasslands. Ecological Applications 9, 45–64.
Tansley, A.G. and Adamson, R.S. (1925) Studies of the vegetation of the English chalk. III. The chalk grasslands of the
Hampshire–Sussex border. Journal of Ecology 13, 177–223.
Taylor, D.R., Aarssen, L.W. and Loehle, C. (1990) On the relationship between r/K selection and environmental carry-
ing capacity: a new habitat templet for plant life history strategies. Oikos 58, 239–250.
Tilman, D. (1982) Resource Competition and Community Structure. Princeton University Press, Princeton, New Jersey,
USA.
Tilman, D. (1987) Secondary succession and the pattern of plant dominance along experimental nitrogen gradients.
Ecological Monographs 57, 189–214.
Tilman, D. (1988) Plant Strategies and the Dynamics and Structure of Plant Communities. Princeton University Press,
Princeton, New Jersey, USA.
Tilman, D. (1990) Constraints and tradeoffs: toward a predictive theory of competition and succession. Oikos 58, 3–15.
Tilman, D. (1997) Community invasibility, recruitment limitation, and grassland biodiversity. Ecology 78, 81–92.
Turkington, R. (1991) Rapid change in a patchy environment – the ‘world’ from a plant’s-eye view. In: Dudley, E.C.
(ed.) The Unity of Biology. Dioscorides, Portland, Oregon, pp. 194–200.
Underwood, A.J. and Petraitis, P.S. (1993) Structure of intertidal assemblages in different locations: how can local
processes be compared. In: Ricklefs, R.E. and Schluter, D. (eds) Species Diversity in Ecological Communities.
University of Chicago Press, Chicago, USA, pp. 39–51.
van der Werf, A., Vannuenen, M., Visser, A.J. and Lambers, H. (1993) Contribution of physiological and morphological
plant traits to a species’ competitive ability at high and low nitrogen supply: a hypothesis for inherently fast-growing
and slow-growing monocotyledonous species. Oecologia 94, 434–440.
Vicari, M. and Bazely, D.R. (1993) Do grasses fight back? The case for antiherbivore defences. Trends in Ecology and
Evolution 8, 137–141.
Vinton, M.A. and Burke, I.C. (1997) Contingent effects of plant species on soils along a regional moisture gradient in
the Great Plains. Oecologia 110, 393–402.
Weaver, J.E. (1950) Effects of different intensities of grazing on depth and quantity of roots of grasses. Journal of Range
Management 3, 100–113.
Wedin, D.A. and Pastor, J. (1993) Nitrogen mineralization dynamics in grass monocultures. Oecologia 96, 186–192.
Wedin, D.A. and Tilman, D. (1993) Competition among grasses along a nitrogen gradient: initial conditions and mecha-
nisms of competition. Ecological Monographs 63, 199–229.
Welden, C.W. and Slauson, W.L. (1986) The intensity of competition versus its importance: an overlooked distinction
and some implications. Quarterly Review of Biology 61, 23–44.
C&S in Pastures Chap 10 22/10/01 1:39 pm Page 212

212 D.A. Peltzer and S.D. Wilson

Werner, P.A. and Platt, W.J. (1976) Ecological relationships of co-occurring goldenrods (Solidago: Compositae).
American Naturalist 110, 959–971.
Westoby, M., Leishman, M. and Lord, J. (1995) Issues of interpretation after relating comparative datasets to phylogeny.
Journal of Ecology 83, 892–893.
Wilson, J.B. and Agnew, A.D.Q. (1992) Positive feedback switches in plant communities. Advances in Ecological
Research 23, 263–336.
Wilson, S.D. (1991) Plasticity, morphology and distribution in twelve lakeshore plants. Oikos 62, 292–298.
Wilson, S.D. (1993a) Competition and resource availability in heath and grassland in the Snowy Mountains of
Australia. Journal of Ecology 81, 445–451.
Wilson, S.D. (1993b) Belowground competition in forest and prairie. Oikos 68, 146–150.
Wilson, S.D. (1994) Initial size and the competitive responses of two grasses at two levels of soil nitrogen: a field experiment.
Canadian Journal of Botany 1349–1354.
Wilson, S.D. (1998) Competition between grasses and woody plants. In: Cheplick, G.P. (ed.) Population Biology of
Grasses. Cambridge University Press, Cambridge, UK, pp. 231–254.
Wilson, S.D. (1999) Plant interactions during secondary succession. In: Walker, L.R. (ed.) Ecosystems of Disturbed
Ground. Elsevier, Amsterdam, The Netherlands, pp. 629–650.
Wilson, S.D. and Keddy, P.A. (1986) Measuring diffuse competition along an environmental gradient: results from a
shoreline plant community. American Naturalist 127, 862–869.
Wilson, S.D. and Shay, J.M. (1990) Competition, fire and nutrients in a mixed-grass prairie. Ecology 71, 1959–1967.
Wilson, S.D. and Tilman, D. (1991) Components of plant competition along an experimental gradient of nitrogen
availability. Ecology 72, 1050–1065.
Wilson, S.D. and Tilman, D. (1993) Plant competition in relation to disturbance, fertility and resource availability.
Ecology 74, 599–611.
Wilson, S.D. and Tilman, D. (1995) Competitive responses of eight old-field plant species in four environments.
Ecology 76, 1169–1180.
Yodzis, P. (1986) Competition, mortality and community structure. In: Diamond, J. and Case, T.J. (eds) Community
Ecology. Harper and Row, New York, USA, pp. 480–491.

You might also like